[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,863)

Search Parameters:
Keywords = biodiesel

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 2549 KiB  
Article
Autocatalytic Acetylation of Crude Glycerol Using Acetic Acid: A Kinetic Model
by Federico M. Perez, Francisco Pompeo, Gerardo F. Santori and Martín N. Gatti
Reactions 2024, 5(3), 472-491; https://doi.org/10.3390/reactions5030025 - 9 Aug 2024
Viewed by 458
Abstract
The aim of this work was to develop a kinetic model based on the power law to describe the evolution of glycerol conversion and product distribution in the crude glycerol (G) acetylation reaction with acetic acid (AA) without the use of a catalyst. [...] Read more.
The aim of this work was to develop a kinetic model based on the power law to describe the evolution of glycerol conversion and product distribution in the crude glycerol (G) acetylation reaction with acetic acid (AA) without the use of a catalyst. For this purpose, experimental tests were carried out with analytical glycerol under different reaction conditions (T = 80–160 °C, AA/G = 1–9 molar ratio, t = 0.25–2 h). The results showed the formation of mono- (MAG), di- (DAG) and tri- (TAG) acetylglycerols, liquid products with multiple applications in the chemical industry. From these results, a kinetic model based on the power law was implemented, which could acceptably estimate the experimental concentrations with an average relative error of 14.9%. Then, crude glycerol samples from different biodiesel industries were characterized by identifying and quantifying the impurities present in them (H2O, CH3OH, NaOH, NaCOOH, MONG and NaCl), and employed as reactants in the reaction tests. Given the significant differences observed in the glycerol conversion values compared to those obtained with analytical glycerol, further reaction tests were conducted to elucidate the effect of each impurity over the glycerol conversion. In these tests, the different impurities were added individually, maintaining the same concentration range as that of the crude glycerol samples. From the results obtained, global activity factors were introduced, which allowed us to modify the kinetic model to estimate glycerol conversions in the crude glycerol samples with an average relative error of 7%. It is hoped that this kinetic model will be a powerful tool useful for designing reactors on an industrial scale. Full article
Show Figures

Figure 1

Figure 1
<p>Activity tests as a function of time using analytical-grade glycerol. (<b>a</b>) Glycerol conversion. (<b>b</b>) Selectivity to MAG. (<b>c</b>) Selectivity to DAG. (<b>d</b>) Selectivity to TAG. Reaction conditions: 80–160 °C, AA/G = 6, P = 2 MPa N<sub>2</sub> initial pressure.</p>
Full article ">Figure 2
<p>Activity tests as a function of the AA/G molar ratio using analytical-grade glycerol. (<b>a</b>) Glycerol conversion. (<b>b</b>) Selectivity to MAG, DAG and TAG. Reaction conditions: 120 °C, 0.5 h, P = 2 MPa N<sub>2</sub> initial pressure.</p>
Full article ">Figure 3
<p>Results of the kinetic model using analytical-grade glycerol. (<b>a</b>) Glycerol concentration as a function of time. (<b>b</b>) MAG concentration as a function of time. (<b>c</b>) DAG concentration as a function of time. (<b>d</b>) TAG concentration as a function of time. (<b>e</b>) Experimental concentration vs. model concentration in a parity plot. Reaction conditions: 80–160 °C, AA/G = 6, P = 2 MPa N<sub>2</sub> initial pressure.</p>
Full article ">Figure 4
<p>Activity tests as a function of time using crude glycerol. (<b>a</b>) Glycerol conversion. (<b>b</b>) Selectivity to MAG. (<b>c</b>) Selectivity to DAG. (<b>d</b>) Selectivity to TAG. Reaction conditions: 80–160 °C, AA/G = 6, P = 2 MPa N<sub>2</sub> initial pressure.</p>
Full article ">Figure 5
<p>Effect of H<sub>2</sub>O addition in the range of 10–50 wt.% over glycerol conversion and selectivity to liquid products. Reaction conditions: 120 °C, AA/G = 6, 0.5 h, P = 2 MPa N<sub>2</sub> initial pressure.</p>
Full article ">Figure 6
<p>Effect of CH<sub>3</sub>OH addition in the range of 5–20 wt.% over glycerol conversion and selectivity to liquid products. Reaction conditions: 120 °C, AA/G = 6, 0.5 h, P = 2 MPa N<sub>2</sub> initial pressure.</p>
Full article ">Figure 7
<p>Effect of neutralization via NaOH addition in the range of 5–30 wt.% over glycerol conversion and selectivity to liquid products. Reaction conditions: 120 °C, AA/G = 6, 0.5 h, P = 2 MPa N<sub>2</sub> initial pressure.</p>
Full article ">Figure 8
<p>Effect of NaCOOH addition in the range of 5–20 wt.% over glycerol conversion and selectivity to liquid products. Reaction conditions: 120 °C, AA/G = 6, 0.5 h, P = 2 MPa N<sub>2</sub> initial pressure.</p>
Full article ">Figure 9
<p>Effect of oleic acid addition in the range of 5–40 wt.% over glycerol conversion and selectivity to liquid products. Reaction conditions: 120 °C, AA/G = 6, 0.5 h, P = 2 MPa N<sub>2</sub> initial pressure.</p>
Full article ">Figure 10
<p>Effect of NaCl addition in the range of 5–20 wt.% over glycerol conversion and selectivity to liquid products. Reaction conditions: 120 °C, AA/G = 6, 0.5 h, P = 2 MPa N<sub>2</sub> initial pressure.</p>
Full article ">Figure 11
<p>(<b>a</b>–<b>e</b>) Glycerol conversion as a function of time for crude glycerol samples A, B, C, D and E. (<b>f</b>) Parity plot of experimental conversion versus model conversion. Reaction conditions: 120 °C, AA/G = 6, P = 2 MPa N2 initial pressure.</p>
Full article ">Figure 11 Cont.
<p>(<b>a</b>–<b>e</b>) Glycerol conversion as a function of time for crude glycerol samples A, B, C, D and E. (<b>f</b>) Parity plot of experimental conversion versus model conversion. Reaction conditions: 120 °C, AA/G = 6, P = 2 MPa N2 initial pressure.</p>
Full article ">Scheme 1
<p>Scheme of reactions involved in glycerol acetylation [<a href="#B8-reactions-05-00025" class="html-bibr">8</a>].</p>
Full article ">Scheme 2
<p>Mechanism of autocatalyzed glycerol acetylation.</p>
Full article ">
23 pages, 2911 KiB  
Review
Reduction and Reuse of Forestry and Agricultural Bio-Waste through Innovative Green Utilization Approaches: A Review
by Jianhui Guo, Yi Zhang, Jianjun Fang, Ziwei Ma, Cheng Li, Mengyao Yan, Naxin Qiao, Yang Liu and Mingming Bian
Forests 2024, 15(8), 1372; https://doi.org/10.3390/f15081372 - 6 Aug 2024
Viewed by 489
Abstract
Biomass waste, which is biodegradable and vastly underutilized, is generated in huge quantities worldwide. Forestry and agricultural biomass wastes are notable for their wide availability, high yield, biodegradability, and recyclability. The accumulation of these wastes not only occupies valuable land but causes serious [...] Read more.
Biomass waste, which is biodegradable and vastly underutilized, is generated in huge quantities worldwide. Forestry and agricultural biomass wastes are notable for their wide availability, high yield, biodegradability, and recyclability. The accumulation of these wastes not only occupies valuable land but causes serious environmental pollution, which can ultimately harm human health. Therefore, leveraging scientific technology to convert forestry and agricultural bio-waste into bioenergy and other valuable products is crucial. In this paper, common forestry and agricultural bio-waste such as straw, rice husks, livestock manure, tree branches, sawdust, and bioenergy (bioethanol, biogas, biodiesel, biohydrogen) were selected as keywords, with the theme of green and efficient utilization. This paper provides a comprehensive review of the sources of biomass waste, existing recycling technologies, and the potential of forestry and agricultural bio-waste as material additives and for conversion to biomass energy and other derivatives, along with future recycling prospects. Full article
Show Figures

Figure 1

Figure 1
<p>Major sources of bio-waste.</p>
Full article ">Figure 2
<p>Environmentally friendly approach for the efficient utilization of biomass waste.</p>
Full article ">Figure 3
<p>Conversion of forestry and agricultural biomass wastes into composite materials.</p>
Full article ">Figure 4
<p>Classification of forestry and agricultural bio-waste available to produce bioenergy [<a href="#B78-forests-15-01372" class="html-bibr">78</a>].</p>
Full article ">Figure 5
<p>Biofuel pathways for obtaining bioethanol from forestry and agricultural bio-waste.</p>
Full article ">
13 pages, 4231 KiB  
Article
On the Influence of Engine Compression Ratio on Diesel Engine Performance and Emission Fueled with Biodiesel Extracted from Waste Cooking Oil
by Jasem Ghanem Alotaibi, Ayedh Eid Alajmi, Talal Alsaeed, Saddam H. Al-Lwayzy and Belal F. Yousif
Energies 2024, 17(15), 3844; https://doi.org/10.3390/en17153844 - 5 Aug 2024
Viewed by 393
Abstract
Despite the extensive research on biodiesels, further investigation is warranted on the impact of compression ratios on emissions and engine performance. This study addresses this gap by evaluating the effects of increasing the engine’s compression ratio on engine performance metrics—brake-specific fuel consumption (BSFC), [...] Read more.
Despite the extensive research on biodiesels, further investigation is warranted on the impact of compression ratios on emissions and engine performance. This study addresses this gap by evaluating the effects of increasing the engine’s compression ratio on engine performance metrics—brake-specific fuel consumption (BSFC), power, torque, and exhaust gas temperature—and emissions—unburnt hydrocarbons (HCs), carbon dioxide (CO2), carbon monoxide (CO), nitrogen oxides (NOx), and oxygen (O2)—when fueled with a 20% blend of waste cooking oil biodiesel (WCB20) and petroleum diesel (PD) under various operating conditions. The viscosity of the prepared fuels was measured at 25 °C and 40 °C. Experiments were conducted on a single-cylinder diesel engine under wide-open throttle conditions at three different speeds (1400 rpm, 2000 rpm, and 2600 rpm) and two compression ratios (16:1 and 18:1). The results revealed that at a lower compression ratio, both WCB20 and petroleum diesel exhibited reduced BSFC compared to higher compression ratios. However, increasing the compression ratio from 16:1 to 18:1 significantly decreased HC emissions but increased CO2 and NOx emissions. Engine power increased with engine speed for both fuels and compression ratios, with WCB20 initially producing less power than diesel but surpassing it at higher compression ratios. WCB20 demonstrated improved combustion quality with lower unburnt hydrocarbons and carbon monoxide emissions due to its higher oxygen content, promoting complete combustion. This study provides critical insights into optimizing engine performance and emission characteristics by manipulating compression ratios and utilizing biodiesel blends, paving the way for more efficient and environmentally friendly diesel engine operations. Full article
(This article belongs to the Topic Advanced Engines Technologies)
Show Figures

Figure 1

Figure 1
<p>Engine setup.</p>
Full article ">Figure 2
<p>The viscosity of the fuels at 25 °C and 40 °C.</p>
Full article ">Figure 3
<p>BSFC (g/kW.h) vs. the engine speed (rpm) at 16:1 and 18:1 CR.</p>
Full article ">Figure 4
<p>Power vs. the engine speed (rpm) at 16:1 and 18:1 CR.</p>
Full article ">Figure 5
<p>Engine torque vs. the engine speed (rpm) at 16:1 and 18:1 CR.</p>
Full article ">Figure 6
<p>Engine exhaust gas temp. vs. the engine speed (rpm) at 16:1 and 18:1 CR.</p>
Full article ">Figure 7
<p>Unburnt hydrocarbon (HC) emission vs. the engine speed (rpm) at 16:1 and 18:1 CR.</p>
Full article ">Figure 8
<p>CO<sub>2</sub> emission vs. the engine speed (rpm) at 16:1 and 18:1 CR.</p>
Full article ">Figure 9
<p>CO emission vs. the engine speed (rpm) at 16:1 and 18:1 CR.</p>
Full article ">Figure 10
<p>Nitrogen oxide (NO<sub>x</sub>) emission vs. the engine speed (rpm) at 16:1 and 18:1 CR.</p>
Full article ">Figure 11
<p>Oxygen gas vs. the engine speed (rpm) at 16:1 and 18:1 CR.</p>
Full article ">
15 pages, 3526 KiB  
Article
Plasma Modification of Biomass-Based Starfish Catalysts for Efficient Biodiesel Synthesis
by Sungho Lee, Jeyoung Ha and Oi Lun Li
Nanomaterials 2024, 14(15), 1313; https://doi.org/10.3390/nano14151313 - 4 Aug 2024
Viewed by 434
Abstract
This study investigated biodiesel production via the transesterification of grapeseed oil with plasma-modified biomass-based catalysts originating from starfish. Dried starfish was first converted into magnesium and calcium oxide through heat treatment and then further modified by plasma engineering to improve the catalyst’s surface [...] Read more.
This study investigated biodiesel production via the transesterification of grapeseed oil with plasma-modified biomass-based catalysts originating from starfish. Dried starfish was first converted into magnesium and calcium oxide through heat treatment and then further modified by plasma engineering to improve the catalyst’s surface area and active sites via zinc addition. The Zn content was added via plasma engineering in the ratios of starfish (Mg0.1Ca0.9CO3): ZnO varying from 5:1, 10:1, to 20:1. The structure and morphology of the catalyst were confirmed through XRD, SEM, and XPS analysis. After the Zn addition and activation process, the surface area and the basicity of the synthesized catalysts were increased. The plasma-modified catalyst showed the highest basicity at the ratio of 10:1. Based on HPLC analyses, the optimized biodiesel yield in transesterification demonstrated 97.7% in fatty acid conversion, and its catalytic performance maintained 93.2% even after three repeated runs. Full article
(This article belongs to the Topic Biomass for Energy, Chemicals and Materials)
Show Figures

Figure 1

Figure 1
<p>The schematic of plasma modification via the plasma process.</p>
Full article ">Figure 2
<p>SEM image of (<b>a</b>) SF, (<b>b</b>) SF700, (<b>c</b>) SF900 (×3k), (<b>d</b>) SFZn1, (<b>e</b>) SFZn2, and (<b>f</b>) SFZn3 (×100k).</p>
Full article ">Figure 3
<p>XRD patterns of the prepared catalysts.</p>
Full article ">Figure 4
<p>XPS spectra of SF900: (<b>a</b>) Ca 2p, (<b>b</b>) Mg 1s, and (<b>c</b>) O 1s.</p>
Full article ">Figure 5
<p>XPS spectra of SFZn2: (<b>a</b>) Ca 2p, (<b>b</b>) Mg 1s, (<b>c</b>) Zn 2p, and (<b>d</b>) O 1s.</p>
Full article ">Figure 6
<p>Possible mechanisms of the transesterification reaction with catalysts.</p>
Full article ">Figure 7
<p>Biodiesel yield of grape seed oil via transesterification under 68 °C: (<b>a</b>) comparison with different reaction times of SFZn2, and (<b>b</b>) transesterification of each sample for 12 h.</p>
Full article ">Figure 8
<p>Biodiesel yield of grape seed oil of each catalyst in three consecutive runs (12 h, 68 °C).</p>
Full article ">
24 pages, 3881 KiB  
Article
Methodological Solutions for Predicting Energy Efficiency of Organic Rankine Cycle Waste Heat Recovery Systems Considering Technological Constraints
by Sergejus Lebedevas and Tomas Čepaitis
J. Mar. Sci. Eng. 2024, 12(8), 1303; https://doi.org/10.3390/jmse12081303 - 1 Aug 2024
Viewed by 450
Abstract
Solving strategic IMO tasks for the decarbonization of maritime transport and the dynamics of its controlling indicators (EEDI, EEXI, CII) involves the comprehensive use of renewable and low-carbon fuels (LNG, biodiesel, methanol in the mid-term perspective of 2030, ammonia, and hydrogen to achieve [...] Read more.
Solving strategic IMO tasks for the decarbonization of maritime transport and the dynamics of its controlling indicators (EEDI, EEXI, CII) involves the comprehensive use of renewable and low-carbon fuels (LNG, biodiesel, methanol in the mid-term perspective of 2030, ammonia, and hydrogen to achieve zero emissions by 2050) and energy-saving technologies. The technology of regenerating secondary heat sources of the ship’s power plant WHR in the form of an Organic Rankine Cycle (ORC) is considered one of the most promising solutions. The attractiveness of the ORC is justified by the share of the energy potential of WHR at 45–50%, almost half of which are low-temperature WHR (80–90 °C and below). However, according to DNV GL, the widespread adoption of WHR-ORC technologies, especially on operating ships, is hindered by the statistical lack of system prototypes combined with the high cost of implementation. Developing methodological tools for justifying the energy efficiency indicators of WHR–ORC cycle implementation is relevant at all stages of design. The methodological solutions proposed in this article are focused on the initial stages of comparative evaluation of alternative structural solutions (without the need to use detailed technical data of the ship’s systems, power plant, and ORC nodes), expected indicators of energy efficiency, and cycle performance. The development is based on generalized results of variation studies of the ORC in the structure of the widely used main marine medium-speed diesel engine Wärtsilä 12V46F (14,400 kW, 500 min−1) in the operational load cycle range of 25–100% of nominal power. The algorithm of the proposed solutions is based on the established interrelationship of the components of the ORC energy balance in the P-h diagram field of thermodynamic indicators of the cycle working fluid (R134a was used). The implemented strategy does allow, in graphical form, for justifying the choice of working fluid and evaluating the energy performance and efficiency of alternative WHR sources for the main engine, taking into account the design solutions of the power turbine and the technological constraints of the ORC condensation system. The verification of the developed methodological solutions is served by the results of comprehensive variation studies of the ORC performed by the authors using the professionally oriented thermoengineering tool “Thermoflow” and the specification data of Wärtsilä 12V46F with an achieved increase in energy efficiency indicators by 21.4–7%. Full article
Show Figures

Figure 1

Figure 1
<p>WHR cycle characteristic points representation in the working material Mollier chart.</p>
Full article ">Figure 2
<p>Algorithmic approach to constructing the WHR cycle configuration and parameter identification. * Error is determined by assessing the accuracy of the initial data of the methodology.</p>
Full article ">Figure 3
<p>Variational estimation scheme of <math display="inline"><semantics> <mrow> <msub> <mi>P</mi> <mrow> <mi>t</mi> <mi>u</mi> <mi>r</mi> <mi>b</mi> </mrow> </msub> <mo>=</mo> <mi>f</mi> <mfenced> <mrow> <msub> <mi>G</mi> <mi>w</mi> </msub> <mo>,</mo> <msub> <mi>π</mi> <mi>T</mi> </msub> <mo>=</mo> <mi>v</mi> <mi>a</mi> <mi>r</mi> </mrow> </mfenced> </mrow> </semantics></math> represented in the Mollier diagram <math display="inline"><semantics> <mrow> <mfenced> <mrow> <msup> <msub> <mi>π</mi> <mi>T</mi> </msub> <mo>′</mo> </msup> <mo>=</mo> <mfrac> <mrow> <msup> <msub> <mi>P</mi> <mn>2</mn> </msub> <mo>′</mo> </msup> </mrow> <mrow> <msub> <mi>P</mi> <mn>1</mn> </msub> </mrow> </mfrac> <mo>;</mo> <mo> </mo> <msup> <msub> <mi>π</mi> <mi>T</mi> </msub> <mrow> <mrow> <mo>″</mo> </mrow> </mrow> </msup> <mo>=</mo> <mfrac> <mrow> <msup> <msub> <mi>P</mi> <mn>2</mn> </msub> <mrow> <mrow> <mo>″</mo> </mrow> </mrow> </msup> </mrow> <mrow> <msub> <mi>P</mi> <mn>1</mn> </msub> </mrow> </mfrac> <mo>;</mo> <mo> </mo> <msup> <msub> <mi>π</mi> <mi>T</mi> </msub> <mrow> <mrow> <mo>‴</mo> </mrow> </mrow> </msup> <mo>=</mo> <mfrac> <mrow> <msup> <msub> <mi>P</mi> <mn>2</mn> </msub> <mrow> <mrow> <mo>‴</mo> </mrow> </mrow> </msup> </mrow> <mrow> <msub> <mi>P</mi> <mn>1</mn> </msub> </mrow> </mfrac> </mrow> </mfenced> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Graphical dependence of the parameters of <math display="inline"><semantics> <mrow> <msub> <mi>P</mi> <mrow> <mi>t</mi> <mi>u</mi> <mi>r</mi> <mi>b</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>G</mi> <mi>w</mi> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <msub> <mi>π</mi> <mi>T</mi> </msub> <mo>=</mo> <mi>v</mi> <mi>a</mi> <mi>r</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>The interrelation of the <math display="inline"><semantics> <mrow> <msub> <mi>P</mi> <mrow> <mi>t</mi> <mi>u</mi> <mi>r</mi> <mi>b</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>G</mi> <mi>w</mi> </msub> </mrow> </semantics></math> parameters with the variants of incorporating possible secondary heat sources in the WHR cycle in a wide load range.</p>
Full article ">Figure A1
<p>ORC–WHR cycle scheme implemented in thermal engineering software “Thermoflow”.</p>
Full article ">
39 pages, 1204 KiB  
Review
Application of Microalgae to Wastewater Bioremediation, with CO2 Biomitigation, Health Product and Biofuel Development, and Environmental Biomonitoring
by Gesthimani Iakovidou, Aikaterini Itziou, Arsenios Tsiotsias, Evangelia Lakioti, Petros Samaras, Constantinos Tsanaktsidis and Vayos Karayannis
Appl. Sci. 2024, 14(15), 6727; https://doi.org/10.3390/app14156727 - 1 Aug 2024
Viewed by 569
Abstract
In the current study, the cultivation of microalgae on wastewater-based substrates is investigated for an effective natural wastewater treatment that also generates biofuels and value-added products beneficial to human health. Additionally, the health of ecosystems can be evaluated via microalgae. The utilization of [...] Read more.
In the current study, the cultivation of microalgae on wastewater-based substrates is investigated for an effective natural wastewater treatment that also generates biofuels and value-added products beneficial to human health. Additionally, the health of ecosystems can be evaluated via microalgae. The utilization of microalgae as bioindicators, biofuel producers, and wastewater treatment providers, under the biorefinery concept, is covered in this article. In fact, bioremediation is feasible, and microalgae culture can be used to efficiently process a variety of effluents. Along with wastewater processing and the creation of value-added substances, bioconversion concurrently offers a viable and promising alternative for reducing CO2 greenhouse gas emissions to contribute to climate change mitigation. The microalgal biorefinery being considered as the third generation is unique in that it addresses all the aforementioned problems, in contrast to lignocellulosic biomass from agricultural waste in second-generation biorefineries and edible crops in first-generation biorefineries. In particular, one of the most promising natural resources for the manufacture of biofuel, including biodiesel, bioethanol, biomethane, and biohydrogen, is found to be microalgae. Furthermore, products of high value, like fatty acid methyl esters, astaxanthin, β-carotene, DHA, and EPA can be made. Hence, microalgal biomass offers a substitute for the development of biofertilizers, bioplastics, pharmaceuticals, cosmetics, animal and aquatic feeds, and human nutrition products, thus promoting human and environmental health. Full article
Show Figures

Figure 1

Figure 1
<p>Biofuel production from microalgae.</p>
Full article ">Figure 2
<p>Simplified flow chart of anaerobic digestion residue treatment with microalgal cultures (modified from Stiles et al. [<a href="#B51-applsci-14-06727" class="html-bibr">51</a>]).</p>
Full article ">Figure 3
<p>Simplified flow diagram of a multi-product microalgal biorefinery (modified from ‘t Lam et al. [<a href="#B53-applsci-14-06727" class="html-bibr">53</a>]).</p>
Full article ">
25 pages, 1430 KiB  
Review
From Citrus Waste to Valuable Resources: A Biorefinery Approach
by Nancy Medina-Herrera, Guillermo Cristian Guadalupe Martínez-Ávila, Claudia Lizeth Robledo-Jiménez, Romeo Rojas and Bianca Sherlyn Orozco-Zamora
Biomass 2024, 4(3), 784-808; https://doi.org/10.3390/biomass4030044 - 1 Aug 2024
Viewed by 492
Abstract
Typically, citrus waste is composted on land by producers or used as livestock feed. However, the biorefinery approach offers a sustainable and economically viable solution for managing and valorizing these agricultural residues. This review examines research from the period 2014 to 2024. Citrus [...] Read more.
Typically, citrus waste is composted on land by producers or used as livestock feed. However, the biorefinery approach offers a sustainable and economically viable solution for managing and valorizing these agricultural residues. This review examines research from the period 2014 to 2024. Citrus waste can be utilized initially by extracting the present phytochemicals and subsequently by producing value-added products using it as a raw material. The phytochemicals reported as extracted include essential oils (primarily limonene), pectin, polyphenolic components, micro- and nano-cellulose, proteins, and enzymes, among others. The components produced from the waste include bioethanol, biogas, volatile acids, biodiesel, microbial enzymes, and levulinic acid, among others. The review indicates that citrus waste has technical, economic, and environmental potential for utilization at the laboratory scale and, in some cases, at the pilot scale. However, research on refining pathways, optimization, and scalability must continue to be an active field of investigation. Full article
Show Figures

Figure 1

Figure 1
<p>Phytochemicals and added-value compounds extracted or produced from citrus by-products. Illustrations were obtained using Microsoft AI image generator v10.</p>
Full article ">Figure 2
<p>Number of papers resulting from the search using the keywords “citrus waste biorefinery” in the Scopus database.</p>
Full article ">
13 pages, 2463 KiB  
Article
Heterogeneous Catalyst Characteristics of TiO2 Nanoparticles Impregnated with Alkaline CH3ONa for Use in Transesterification Process
by Cherng-Yuan Lin and Shun-Lien Tseng
Processes 2024, 12(8), 1584; https://doi.org/10.3390/pr12081584 - 29 Jul 2024
Viewed by 380
Abstract
A strong alkaline catalyst, sodium methoxide (CH3ONa), is commonly used to catalyze the transesterification reaction for biodiesel production. Meanwhile, titanium dioxide (TiO2) anatase with a bandgap of 3.2 eV is a highly competitive photocatalyst after the absorption of sufficient [...] Read more.
A strong alkaline catalyst, sodium methoxide (CH3ONa), is commonly used to catalyze the transesterification reaction for biodiesel production. Meanwhile, titanium dioxide (TiO2) anatase with a bandgap of 3.2 eV is a highly competitive photocatalyst after the absorption of sufficient energy from ultraviolet light. There has been no published report on the synergistic catalyst effects of CH3ONa and TiO2 on further facilitating the transesterification reaction. Hence, an impregnating method was used in this study to prepare the heterogeneous photocatalyst comprising TiO2 nanoparticles embedded with a CH3ONa catalyst. The TiO2 nanoparticles were first immersed in an aqueous solution of CH3ONa so that CH3ONa could diffuse into the interior surfaces of the TiO2 porous structure. The mixture of TiO2 and CH3ONa was then calcined in the temperature range from 150 °C to 450 °C for 4 h to produce the TiO2/CH3ONa photocatalyst. Various characteristics of the catalyst were analyzed to determine the optimum preparation conditions. The Fourier transform infrared spectroscopy spectra revealed that the absorption peaks of CH3ONa appeared in the wavelength range of 600 cm−1 and 1500 cm−1. The X-ray diffractometer analysis showed that the calcined CH3ONa did not alter the crystal structure of the catalyst carrier TiO2. At the calcined temperatures between 100 °C and 800 °C, no intermediate or pyrolyzed product of CH3ONa was detected, as revealed by the thermogravimetric analyzer spectra. In addition, about 5~9 wt.% elemental calcium in the CH3ONa solution could be calcined onto the surface of TiO2. In addition, the FTIR spectra confirmed the successful sintering and bonding of CH3ONa onto the TiO2 nanoparticles. The energy dispersive spectrometry result revealed that the interior surface of the TiO2 nanoparticles was filled with the CH3ONa compound. Full article
(This article belongs to the Section Environmental and Green Processes)
Show Figures

Figure 1

Figure 1
<p>Preparation procedures for the synthesized photocatalyst TiO<sub>2</sub>/CH<sub>3</sub>ONa.</p>
Full article ">Figure 2
<p>The temperature-control program for the calcined catalyst in the high-temperature furnace.</p>
Full article ">Figure 3
<p>Spectra of (<b>a</b>) catalyst components of sodium methoxide (CH<sub>3</sub>Ona), (<b>b</b>) titanium dioxide (TiO<sub>2</sub>) powder, (<b>c</b>) solid alkaline catalyst TiO<sub>2</sub>/CH<sub>3</sub>Ona sintered at 200 °C, and (<b>d</b>) solid alkaline catalyst TiO<sub>2</sub>/CH<sub>3</sub>Ona sintered at temperatures between 150 °C and 450 °C analyzed by Fourier transform infrared spectroscopy.</p>
Full article ">Figure 3 Cont.
<p>Spectra of (<b>a</b>) catalyst components of sodium methoxide (CH<sub>3</sub>Ona), (<b>b</b>) titanium dioxide (TiO<sub>2</sub>) powder, (<b>c</b>) solid alkaline catalyst TiO<sub>2</sub>/CH<sub>3</sub>Ona sintered at 200 °C, and (<b>d</b>) solid alkaline catalyst TiO<sub>2</sub>/CH<sub>3</sub>Ona sintered at temperatures between 150 °C and 450 °C analyzed by Fourier transform infrared spectroscopy.</p>
Full article ">Figure 4
<p>Thermogravimetric analysis of CH<sub>3</sub>Ona at the temperature range of 40 °C to 1000 °C.</p>
Full article ">Figure 5
<p>X-ray intensity spectra of pure TiO<sub>2</sub> and TiO<sub>2</sub>/CH<sub>3</sub>Ona calcined at 150 °C~450 °C.</p>
Full article ">Figure 6
<p>Photos of scanning electron microscopy of the TiO<sub>2</sub>/CH<sub>3</sub>Ona catalyst under various forging temperatures from 150 °C to 450 °C at (<b>a</b>) 150 °C, (<b>b</b>) 200 °C, (<b>c</b>) 250 °C, (<b>d</b>) 300 °C, (<b>e</b>) 350 °C, and (<b>f</b>) 450 °C forging temperatures.</p>
Full article ">Figure 6 Cont.
<p>Photos of scanning electron microscopy of the TiO<sub>2</sub>/CH<sub>3</sub>Ona catalyst under various forging temperatures from 150 °C to 450 °C at (<b>a</b>) 150 °C, (<b>b</b>) 200 °C, (<b>c</b>) 250 °C, (<b>d</b>) 300 °C, (<b>e</b>) 350 °C, and (<b>f</b>) 450 °C forging temperatures.</p>
Full article ">
16 pages, 3041 KiB  
Article
An Evaluation of the Effect of Fuel Injection on the Performance and Emission Characteristics of a Diesel Engine Fueled with Plastic-Oil–Hydrogen–Diesel Blends
by Kodandapuram Jayasimha Reddy, Gaddale Amba Prasad Rao, Reddygari Meenakshi Reddy and Upendra Rajak
Appl. Sci. 2024, 14(15), 6539; https://doi.org/10.3390/app14156539 - 26 Jul 2024
Viewed by 393
Abstract
Fuelled engines serve as prime movers in low-, medium-, and heavy-duty applications with high thermal diesel efficiency and good fuel economy compared to their counterpart, spark ignition engines. In recent years, diesel engines have undergone a multitude of developments, however, diesel engines release [...] Read more.
Fuelled engines serve as prime movers in low-, medium-, and heavy-duty applications with high thermal diesel efficiency and good fuel economy compared to their counterpart, spark ignition engines. In recent years, diesel engines have undergone a multitude of developments, however, diesel engines release high levels of NOx, smoke, carbon monoxide [CO], and hydrocarbon [HC] emissions. Due to the exponential growth in fleet population, there is a severe burden caused by petroleum-derived fuels. To tackle both fuel and pollution issues, the research community has developed strategies to use economically viable alternative fuels. The present experimental investigations deal with the use of blends of biodiesel prepared from waste plastic oil [P] and petro-diesel [D], and, to improve its performance, hydrogen [H] is added in small amounts. Further, advanced injection timings have been adopted [17.5° to 25.5° b TDC (before top dead centre)] to study their effect on harmful emissions. Hydrogen energy shares vary from 5 to 15%, maintaining a biodiesel proportion of 20%, and the remaining is petro-diesel. Thus, the adopted blends are DP20 ((diesel fuel (80%) and waste plastic biofuel (20%)), DP20H5 (DP20 (95%) and hydrogen (5%)), DP20H10 (DP20 (90%) and hydrogen (10%)), and DP20H15 (DP20 (85%) and hydrogen (15%)). The experiments were conducted at constant speeds with a rated injection pressure of 220 bar and a rated compression ratio of 18. The increase in the share of hydrogen led to a considerable improvement in the performance. Under full load conditions, with advanced injection timings, the brake-specific fuel consumption had significantly decreased and NOx emissions increased. Full article
Show Figures

Figure 1

Figure 1
<p>Blend fuel and diesel fuel.</p>
Full article ">Figure 2
<p>(<b>a</b>) Pictorial view of experimental setup, (<b>b</b>) Schematic diagram of experimental setup.</p>
Full article ">Figure 3
<p>BTE (%) with injection timing for hydrogen blends.</p>
Full article ">Figure 4
<p>SFC with injection timing for hydrogen blends.</p>
Full article ">Figure 5
<p>EGT (°C) with injection timing for hydrogen blends.</p>
Full article ">Figure 6
<p>Cylinder pressure with injection timing for hydrogen blends.</p>
Full article ">Figure 7
<p>MRPR with injection timing and hydrogen blends.</p>
Full article ">Figure 8
<p>ID with injection timing and hydrogen blends.</p>
Full article ">Figure 9
<p>Smoke emissions with injection timing and hydrogen blends.</p>
Full article ">Figure 10
<p>NOx emissions with injection timing and hydrogen blends.</p>
Full article ">Figure 11
<p>CO emissions with injection timing and hydrogen blends.</p>
Full article ">Figure 12
<p>HC emissions with injection timing and hydrogen blends.</p>
Full article ">
15 pages, 3639 KiB  
Article
Biodiesel Production from Waste Cooking Oil Using Recombinant Escherichia coli Cells Immobilized into Fe3O4–Chitosan Magnetic Microspheres
by Zexin Zhao, Meiling Han, Ling Zhou, Changgao Wang, Jianguo Lin, Xin Du and Jun Cai
Molecules 2024, 29(15), 3469; https://doi.org/10.3390/molecules29153469 - 24 Jul 2024
Viewed by 425
Abstract
Developing reusable and easy-to-operate biocatalysts is of significant interest in biodiesel production. Here, magnetic whole-cell catalysts constructed through immobilizing recombinant Escherichia coli cells (containing MAS1 lipase) into Fe3O4–chitosan magnetic microspheres (termed MWCC@MAS1) were used for fatty acid methyl ester [...] Read more.
Developing reusable and easy-to-operate biocatalysts is of significant interest in biodiesel production. Here, magnetic whole-cell catalysts constructed through immobilizing recombinant Escherichia coli cells (containing MAS1 lipase) into Fe3O4–chitosan magnetic microspheres (termed MWCC@MAS1) were used for fatty acid methyl ester (FAME) production from waste cooking oil (WCO). During the preparation process of immobilized cells, the effects of chitosan concentration and cell concentration on their activity and activity recovery were investigated. Optimal immobilization was achieved with 3% (w/v) chitosan solution and 10 mg wet cell/mL cell suspension. Magnetic immobilization endowed the whole-cell catalysts with superparamagnetism and improved their methanol tolerance, enhancing the recyclability of the biocatalysts. Additionally, we studied the effects of catalyst loading, water content, methanol content, and reaction temperature on FAME yield, optimizing these parameters using response surface methodology and Box–Behnken design. An experimental FAME yield of 89.19% was gained under the optimized conditions (3.9 wt% catalyst loading, 22.3% (v/w) water content, 23.0% (v/w) methanol content, and 32 °C) for 48 h. MWCC@MAS1 demonstrated superior recyclability compared to its whole-cell form, maintaining about 86% of its initial productivity after 10 cycles, whereas the whole-cell form lost nearly half after just five cycles. These results suggest that MWCC@MAS1 has great potential for the industrial production of biodiesel. Full article
(This article belongs to the Section Green Chemistry)
Show Figures

Figure 1

Figure 1
<p>Optimization of the process of immobilizing recombinant <span class="html-italic">E. coli</span> cells in magnetic chitosan microspheres: (<b>A</b>) Effect of chitosan concentration on lipase activity and activity recovery of MWCC@MAS1. (<b>B</b>) Effect of cell concentration on lipase activity and activity recovery of MWCC@MAS1.</p>
Full article ">Figure 2
<p>Characterization of the developed MWCC@MAS1 system: (<b>A</b>) Appearance of the magnetic particles. (<b>B</b>) Micromorphology of MWCC@MAS1 obversed with SEM. (<b>C</b>) Comparison of the XRD patterns between MWCC@MAS1 (black line) and Fe<sub>3</sub>O<sub>4</sub> (red line). (<b>D</b>) Magnetization curves of MWCC@MAS1 at room temperature.</p>
Full article ">Figure 3
<p>Single factor optimization of biodiesel production conditions catalyzed by MWCC@MAS1: (<b>A</b>) Effect of catalyst loading on the reaction equilibrium. Reaction systems contained 2 g WCO, 30% (<span class="html-italic">v</span>/<span class="html-italic">w</span>) water content, 30% (<span class="html-italic">v</span>/<span class="html-italic">w</span>) methanol content, and a certain amount of MWCC@MAS1. (<b>B</b>) Effect of water content on FAME yield. Reaction systems contained 2 g WCO, 3 wt% magnetic catalysts, 30% (<span class="html-italic">v</span>/<span class="html-italic">w</span>) methanol content, and a certain volume of reaction buffer. (<b>C</b>) Effect of methanol content on FAME yield. Reaction systems contained 2 g WCO, 25% (<span class="html-italic">v</span>/<span class="html-italic">w</span>) water content, 3 wt% magnetic catalysts, and a certain volume of methanol. Reactions mentioned in (<b>A</b>–<b>C</b>) were carried out at 30 °C and 400 rpm magnetic stirring, and the reaction time of (<b>B</b>,<b>C</b>) was 48 h. (<b>D</b>) Effect of temperature on FAME yield. Reaction systems contained 2 g WCO, 3 wt% magnetic catalysts, 25% (<span class="html-italic">v</span>/<span class="html-italic">w</span>) water content, and 25% (<span class="html-italic">v</span>/<span class="html-italic">w</span>) methanol content, and the reactions were performed at a certain temperature and 400 rpm stirring for 48 h.</p>
Full article ">Figure 4
<p>Three-dimensional response surface plots of variables influencing FAME yield from WCO: (<b>A</b>) Effect of catalyst loading and water content on FAME yield. (<b>B</b>) Effect of catalyst loading and methanol content on FAME yield. (<b>C</b>) Effect of temperature and catalyst loading on FAME yield. (<b>D</b>) Effect of methanol content and water content on FAME yield. (<b>E</b>) Effect of water content and temperature on FAME yield. (<b>F</b>) Effect of methanol content and temperature on FAME yield.</p>
Full article ">Figure 5
<p>Fatty acid composition analysis of WCO and biodiesel product.</p>
Full article ">Figure 6
<p>Recycling of the WCC@MAS1 and MWCC@MAS1 in the conversion of WCO into FAME. The WCC-catalyzed reaction system contained 2 g WCO, 1.5 wt% WCC@MAS1, 30% (<span class="html-italic">v</span>/<span class="html-italic">w</span>) reaction buffer and 30% (<span class="html-italic">v</span>/<span class="html-italic">w</span>) methanol. The reactions were performed at 29 °C and 400 rpm magnetic stirring for 48 h. T WCC-catalyzed reaction system contained 2 g WCO, 3.9 wt% catalyst loading, 22.3% (<span class="html-italic">v</span>/<span class="html-italic">w</span>) water content, 23.0% (<span class="html-italic">v</span>/<span class="html-italic">w</span>) methanol content, and 32 °C) were used in MWCC-catalyzed reaction system. The FAME yield in the first cycle was set as 100% productivity.</p>
Full article ">
15 pages, 1945 KiB  
Review
A Review of Grease Trap Waste Management in the US and the Upcycle as Feedstocks for Alternative Diesel Fuels
by Andres Mata, Junsong Zhang, Joshua Pridemore, Kevin Johnson, Nathan Holliday, Art Helmstetter and Mingming Lu
Environments 2024, 11(8), 159; https://doi.org/10.3390/environments11080159 - 23 Jul 2024
Viewed by 664
Abstract
As byproducts generated by commercial and domestic food-related processes, FOGs (fats, oils, and grease) are the leading cause of sewer pipe blockages in the US and around the world. Grease trap waste (GTW) is a subcategory of FOG currently disposed of as waste, [...] Read more.
As byproducts generated by commercial and domestic food-related processes, FOGs (fats, oils, and grease) are the leading cause of sewer pipe blockages in the US and around the world. Grease trap waste (GTW) is a subcategory of FOG currently disposed of as waste, resulting in an economic burden for GTW generators and handlers. This presents a global need for both resource conservation and carbon footprint reduction, particularly through increased waste upcycling. Therefore, it is critical to better understand current GTW handling practices in the context of the urban food–energy–water cycle. This can be accomplished with firsthand data collection, such as onsite visits, phone discussions, and targeted questionnaires. GTW disposal methods were found to be regional and correspond to key geographical locations, with landfill operations mostly practiced in the Midwest regions, incineration mainly in the Northeast and Mid-Atlantic regions, and digestion mainly in the West of the US. Select GTW samples were analyzed to evaluate their potential reuse as low-cost feedstocks for biodiesel or renewable diesel, which are alternatives to petroleum diesel fuels. Various GTW lipid extraction technologies have been reviewed, and more studies were found on converting GTW into biodiesel rather than renewable diesel. The challenges for these two pathways are the high sulfur content in biodiesel and the metal contents in renewable diesel, respectively. GTW lipid extraction technologies should overcome these issues while producing minimum-viable products with higher market values. Full article
(This article belongs to the Special Issue Environments: 10 Years of Science Together)
Show Figures

Figure 1

Figure 1
<p>Generation and upcycle of oil-containing byproducts from food preparation processes.</p>
Full article ">Figure 2
<p>Locations of all publicly owned treatment works (POTWs) interviewed (POTWs in blue are among the 100 largest POTWs).</p>
Full article ">Figure 3
<p>GTW disposal in open dumpsters (<b>a</b>) and in filter bags (<b>b</b>) at a POTW in southeastern Ohio (OHTW).</p>
Full article ">Figure 4
<p>Differences between biodiesel and renewable diesel from oil feedstocks.</p>
Full article ">Figure 5
<p>Potential upcycle pathways of GTW in urban food–energy–water systems.</p>
Full article ">
17 pages, 2923 KiB  
Article
Comparison of Scenedesmus obliquus in CO2 Capture, Biolipid Production and Nutrient Removal
by Wenwen Cao, Hongfei Yu, Wei Dong, Zijia Rong, Dianbao Peng, Fukun Chen and Lixin Li
Separations 2024, 11(7), 218; https://doi.org/10.3390/separations11070218 - 22 Jul 2024
Viewed by 693
Abstract
The cultivation of microalgae from municipal wastewater, while simultaneously removing nutrients from the water column, has the potential to aid biodiesel production and carbon dioxide fixation, thereby alleviating the pressure of energy shortages. In this research, different ratios of sodium bicarbonate and glucose [...] Read more.
The cultivation of microalgae from municipal wastewater, while simultaneously removing nutrients from the water column, has the potential to aid biodiesel production and carbon dioxide fixation, thereby alleviating the pressure of energy shortages. In this research, different ratios of sodium bicarbonate and glucose were used to prepare simulated municipal wastewater. The results demonstrated that microalgae were most effectively treated under one-stage direct treatment conditions. During direct culture, the most effective treatment was observed for IAA-3, which exhibited a dry weight of 1.4363 g/L and a lipid content of 25.05% after stimulation with 0.0005 M NaHCO3. In contrast, NaHCO3-2 demonstrated optimal performance during the secondary culture, with a dry weight of 1.6844 g/L and a lipid content of 18.05%. Finally, the economic, social and environmental benefits of direct treatment (IAA-3) and secondary treatment NaHCO3-2 were analyzed. The benefits of direct treatment were found to be USD 0.50989/L, while those of secondary treatment were USD 0.43172/L. For each tonne of municipal wastewater treated, the carbon sequestration benefits of IAA-3 during direct treatment and NaHCO3-2 during secondary treatment were USD 0.45645 and USD 0.85725, respectively. Full article
Show Figures

Figure 1

Figure 1
<p>Dry weight (DW) of microalgal biomass at different NaHCO<sub>3</sub> dosing ratios.</p>
Full article ">Figure 2
<p>Lipid production and lipid content of microalgae under different NaHCO<sub>3</sub> dosing ratios.</p>
Full article ">Figure 3
<p>Nutrient removal effects at different NaHCO<sub>3</sub> dosing rates.</p>
Full article ">Figure 4
<p>Dry weight (<b>a</b>), lipid production (<b>b</b>), and decontamination (<b>c</b>) of microalgae cultured directly for the first 7 days.</p>
Full article ">Figure 5
<p>Dry weight (<b>a</b>), lipid production (<b>b</b>), and removal efficiency (<b>c</b>) of microalgae cultured directly for the post 7 days.</p>
Full article ">Figure 6
<p>Dry weight (<b>a</b>), lipid production (<b>b</b>), and removal efficiency (<b>c</b>) of microalgae in secondary cultured for the post 7 days.</p>
Full article ">
14 pages, 6031 KiB  
Article
Reaction Mechanism of Pyrolysis and Combustion of Methyl Oleate: A ReaxFF-MD Analysis
by Yu Wei, Xiaohui Zhang, Shan Qing and Hua Wang
Energies 2024, 17(14), 3536; https://doi.org/10.3390/en17143536 - 18 Jul 2024
Viewed by 349
Abstract
As an emerging environmentally friendly fuel, biodiesel has excellent fuel properties comparable to those of petrochemical diesel. Oleic acid methyl ester, as the main component of biodiesel, has the characteristics of high cetane number and low emission rate of harmful gases. However, the [...] Read more.
As an emerging environmentally friendly fuel, biodiesel has excellent fuel properties comparable to those of petrochemical diesel. Oleic acid methyl ester, as the main component of biodiesel, has the characteristics of high cetane number and low emission rate of harmful gases. However, the comprehensive chemical conversion pathway of oleic acid methyl ester is not clear. In this paper, the reactive force field molecular dynamics simulation (ReaxFF-MD) method is used to construct a model of oleic acid methyl ester pyrolysis and combustion system. Further, the chemical conversion kinetics process at high temperatures (2500 K–3500 K) was studied, and a chemical reaction network was drawn. The research results show that the density of the system has almost no effect on the decomposition activation energy of oleic acid methyl ester, and the activation energies of its pyrolysis and combustion processes are 190.02 kJ/mol and 144.89 kJ/mol, respectively. Ethylene, water and carbon dioxide are the dominant and most accumulated products. From the specific reaction mechanism, the main pyrolysis path of oleic acid methyl ester is the breakage of the C-C bond to produce small molecule intermediates, and subsequent transformation of the ester group radical into carbon oxides. The combustion path is the evolution of long-chain alkanes into short-carbon-chain gaseous products, and these species are further burned to form stable CO2 and H2O. This study further discusses the microscopic combustion kinetics of biodiesel, providing a reference for the construction of biodiesel combustion models. Based on this theoretical study, the understanding of free radicals, intermediates, and products in the pyrolysis and combustion of biomass can be deepened. Full article
(This article belongs to the Section I2: Energy and Combustion Science)
Show Figures

Figure 1

Figure 1
<p>Molecular models of C<sub>19</sub>H<sub>36</sub>O<sub>2</sub> and C<sub>19</sub>H<sub>36</sub>O<sub>2</sub>/O<sub>2</sub>.</p>
Full article ">Figure 2
<p>Quantitative changes of C<sub>19</sub>H<sub>36</sub>O<sub>2</sub> at different system densities.</p>
Full article ">Figure 3
<p>Quantitative changes in the combustion of C<sub>19</sub>H<sub>36</sub>O<sub>2</sub> at different system densities.</p>
Full article ">Figure 4
<p>Changes in number of C<sub>19</sub>H<sub>36</sub>O<sub>2</sub> during the reaction: (<b>a</b>) pyrolysis and (<b>c</b>) combustion, and decomposition times of C<sub>19</sub>H<sub>36</sub>O<sub>2</sub> (<b>b</b>) pyrolysis and (<b>d</b>) combustion.</p>
Full article ">Figure 5
<p>Fitting relationship between the reaction rate and temperature of C<sub>19</sub>H<sub>36</sub>O<sub>2</sub>.</p>
Full article ">Figure 6
<p>Main pyrolysis products at different temperatures: (<b>a</b>) 2000 K, (<b>b</b>) 2500 K, (<b>c</b>) 3000 K and (<b>d</b>) 3500 K.</p>
Full article ">Figure 7
<p>Main combustion products at different temperatures: (<b>a</b>) 2000 K, (<b>b</b>) 2500 K, (<b>c</b>) 3000 K and (<b>d</b>) 3500 K.</p>
Full article ">Figure 8
<p>C<sub>2</sub>H<sub>4</sub> product distribution at different temperatures: (<b>a</b>) pyrolysis and (<b>b</b>) combustion.</p>
Full article ">Figure 9
<p>H<sub>2</sub> product distribution at different temperatures: (<b>a</b>) pyrolysis and (<b>b</b>) combustion.</p>
Full article ">Figure 10
<p>Distribution of major oxidation products of C<sub>19</sub>H<sub>36</sub>O<sub>2</sub> combustion: (<b>a</b>) H<sub>2</sub>O, (<b>b</b>) CH<sub>2</sub>O, (<b>c</b>) CO<sub>2</sub> and (<b>d</b>) CO.</p>
Full article ">Figure 11
<p>Main product reaction pathways: (<b>a</b>) C<sub>2</sub>H<sub>4</sub>, (<b>b</b>) CH<sub>4</sub>, (<b>c</b>) H<sub>2</sub>O, (<b>d</b>) CH<sub>2</sub>O, (<b>e</b>) CO<sub>2</sub> and (<b>f</b>) CO.</p>
Full article ">Figure 12
<p>Reaction mechanism of C<sub>19</sub>H<sub>36</sub>O<sub>2</sub>: (<b>a</b>) Pyrolysis, (<b>b</b>) Combustion.</p>
Full article ">
14 pages, 5642 KiB  
Article
From Marginal Lands to Biofuel Bounty: Predicting the Distribution of Oilseed Crop Idesia polycarpa in Southern China’s Karst Ecosystem
by Yangyang Wu, Panli Yuan, Siliang Li, Chunzi Guo, Fujun Yue, Guangjie Luo, Xiaodong Yang, Zhonghua Zhang, Ying Zhang, Jinli Yang, Haobiao Wu and Guanghong Zhou
Agronomy 2024, 14(7), 1563; https://doi.org/10.3390/agronomy14071563 - 18 Jul 2024
Viewed by 439
Abstract
With the global energy crisis and the decline of fossil fuel resources, biofuels are gaining attention as alternative energy sources. China, as a major developing country, has long depended on coal and is now looking to biofuels to diversify its energy structure and [...] Read more.
With the global energy crisis and the decline of fossil fuel resources, biofuels are gaining attention as alternative energy sources. China, as a major developing country, has long depended on coal and is now looking to biofuels to diversify its energy structure and ensure sustainable development. However, due to its large population and limited arable land, it cannot widely use corn or sugarcane as raw materials for bioenergy. Instead, the Chinese government encourages the planting of non-food crops on marginal lands to safeguard food security and support the biofuel sector. The Southern China Karst Region, with its typical karst landscape and fragile ecological environment, offers a wealth of potential marginal land resources that are suitable for planting non-food energy crops. This area is also one of the most impoverished rural regions in China, confronting a variety of challenges, such as harsh natural conditions, scarcity of land, and ecological deterioration. Idesia polycarpa, as a fast-growing tree species that is drought-tolerant and can thrive in poor soil, is well adapted to the karst region and has important value for ecological restoration and biodiesel production. By integrating 19 bioclimatic variables and karst landform data, our analysis reveals that the Maximum Entropy (MaxEnt) model surpasses the Random Forest (RF) model in predictive accuracy for Idesia polycarpa’s distribution. The karst areas of Sichuan, Chongqing, Hubei, Hunan, and Guizhou provinces are identified as highly suitable for the species, aligning with regions of ecological vulnerability and poverty. This research provides critical insights into the strategic cultivation of Idesia polycarpa, contributing to ecological restoration, local economic development, and the advancement of China’s biofuel industry. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>The study area of the Southern Karst Region of China. YGGKDR (Yunnan–Guizhou–Guangxi Karst Desertification Region), WYBMR (Western Yunnan Border Mountain Region), FPTR (Four Provinces Tibetan Region), LXMR (Luoxiao Mountain Region), WLMR (Wuling Mountain Region), DBMR (Dabie Mountain Region), WMMR (Wumeng Mountain Region), and QBMR (Qinba Mountain Region).</p>
Full article ">Figure 2
<p>Heatmap of the correlation analysis between environmental variables.</p>
Full article ">Figure 3
<p>Accuracy of species distribution model forecasts: (<b>a</b>) Maximum Entropy model; (<b>b</b>) Random Forest model.</p>
Full article ">Figure 4
<p>Spatial distribution and provincial proportion of suitable areas for <span class="html-italic">Idesia polycarpa</span> in the Southern China Karst Region: (<b>a</b>) spatial distribution of suitable areas for <span class="html-italic">Idesia polycarpa</span> in the Southern China Karst Region; (<b>b</b>) proportion of suitable areas for <span class="html-italic">Idesia polycarpa</span> in each province of the Southern China Karst Region.</p>
Full article ">Figure 5
<p>The proportion and area of suitable areas of karst and non-karst in karst areas of Southern China: (<b>a</b>) proportion of suitable areas of karst and non-karst at all levels; (<b>b</b>) karst and non-karst suitable area of each grade.</p>
Full article ">Figure 6
<p>The distribution and proportion of suitable areas of each grade in contiguous poverty-stricken areas: (<b>a</b>) distribution of suitable areas of each grade in contiguous poverty-stricken areas; (<b>b</b>) proportion of suitable areas of each grade in contiguous poverty-stricken areas. YGGKDR (Yunnan–Guizhou–Guangxi Karst Desertification Region), WYBMR (Western Yunnan Border Mountain Region), FPTR (Four Provinces Tibetan Region), LXMR (Luoxiao Mountain Region), WLMR (Wuling Mountain Region), DBMR (Dabie Mountain Region), WMMR (Wumeng Mountain Region), and QBMR (Qinba Mountain Region).</p>
Full article ">
25 pages, 544 KiB  
Article
A Comprehensive Approach to Biodiesel Blend Selection Using GRA-TOPSIS: A Case Study of Waste Cooking Oils in Egypt
by Marwa M. Sleem, Osama Y. Abdelfattah, Amr A. Abohany and Shaymaa E. Sorour
Sustainability 2024, 16(14), 6124; https://doi.org/10.3390/su16146124 - 17 Jul 2024
Viewed by 589
Abstract
The transition to sustainable energy sources is critical for addressing global environmental challenges. In 2017, Egypt produced about 500,000 tons of waste cooking oil from various sources including food industries, restaurants and hotels. Sadly, 90% of households choose to dispose of their used [...] Read more.
The transition to sustainable energy sources is critical for addressing global environmental challenges. In 2017, Egypt produced about 500,000 tons of waste cooking oil from various sources including food industries, restaurants and hotels. Sadly, 90% of households choose to dispose of their used cooking oil by pouring it down the drain or into their village’s sewers instead of using proper disposal methods. The process involves converting waste cooking oil (WCO) into biodiesel.This study introduces a multi-criteria decision-making approach to identify the optimal biodiesel blend from waste cooking oils in Egypt. By leveraging the grey relational analysis (GRA) combined with the technique for order preference by similarity to the ideal solution (TOPSIS), we evaluate eight biodiesel blends (diesel, B5, B10, B20, B30, B50, B75, B100) against various performance metrics, including carbon monoxide, carbon dioxide, nitrogen oxides, hydrocarbons, particulate matter, engine power, fuel consumption, engine noise, and exhaust gas temperature. The experimental analysis used a single-cylinder, constant-speed, direct-injection eight cylinder diesel engine under varying load conditions. Our methodology involved feature engineering and model building to enhance predictive accuracy. The results demonstrated significant improvements in monitoring accuracy, with diesel, B5, and B20 emerging as the top-performing blends. Notably, the B5 blend showed the best overall performance, balancing efficiency and emissions. This study highlights the potential of integrating advanced AI-driven decision-making frameworks into biodiesel blend selection, promoting cleaner energy solutions and optimizing engine performance. Our findings underscore the substantial benefits of waste cooking oils for biodiesel production, contributing to environmental sustainability and energy efficiency. Full article
(This article belongs to the Special Issue Sustainable Materials, Manufacturing and Design)
Show Figures

Figure 1

Figure 1
<p>Transesterification process.</p>
Full article ">Figure 2
<p>Decision hierarchy.</p>
Full article ">Figure 3
<p>Schematic of the complete test rig.</p>
Full article ">Figure 4
<p>Photograph of SLM.</p>
Full article ">
Back to TopTop